Using computer simulations, we validate a simple free energy model that can be analytically solved to predict the equilibrium size of self-limiting clusters of particles in the fluid state governed by a combination of short-range attractive and long-range repulsive pair potentials. The model is a semi-empirical adaptation and extension of the canonical free energy-based result due to Groenewold and Kegel [J. Phys. Chem. B 105, 11702–11709 (2001)], where we use new computer simulation data to systematically improve the cluster-size scalings with respect to the strengths of the competing interactions driving aggregation. We find that one can adapt a classical nucleation like theory for small energetically frustrated aggregates provided one appropriately accounts for a size-dependent, microscopic energy penalty of interface formation, which requires new scaling arguments. This framework is verified in part by considering the extensive scaling of intracluster bonding, where we uncover a superlinear scaling regime distinct from (and located between) the known regimes for small and large aggregates. We validate our model based on comparisons against approximately 100 different simulated systems comprising compact spherical aggregates with characteristic (terminal) sizes between six and sixty monomers, which correspond to wide ranges in experimentally controllable parameters.

Over the past century, colloidal aggregation has been observed and described in a wide range of contexts via progressively more powerful experimental techniques, phenomenological frameworks, and quantitative models.1–6 Spanning processes including droplet nucleation and growth, gel and glass formation, various self-assembly processes, etc., an overarching goal has been to use statistical mechanical or molecular thermodynamic approaches adopted from atomic systems and, if necessary, empirical rules to relate the strength and lengthscale of particle interactions to resulting equilibrium (or non-equilibrium) structures and the thermodynamics (and kinetics) of their formation. These types of relations, especially when based on physically intuitive thermodynamic arguments, are not only of fundamental importance but also highlight pathways for engineering materials at the nano- to microscopic level.

In this article, we focus on fluids where interactions between primary particles (monomers) are characterized by attractions acting at small lengthscales close to contact that compete with repulsions acting at larger lengthscales. This class of interactions can drive the reversible formation of equilibrium cluster phases composed of self-terminating aggregates (droplets) of monomers. Such cluster phases have been the focus of much recent work, ranging from theoretical and computational studies of idealized colloidal or nanoparticle suspensions7–19 to experimental demonstrations for both archetypal colloidal particles20–23 and heterogeneous monomers with anisotropic interactions like proteins.24–31 Despite the range of materials and lengthscales, the broad underlying physics appear universal: induce (or allow) depletion (dispersion) attractions between monomers to drive aggregation while simultaneously controlling electrostatic repulsions between the ionic double-layers of monomers such that they collectively build up to attenuate growth.

While this basic paradigm of frustrating interactions is well-accepted, it is not yet established how to best describe observed cluster phases in terms of their thermodynamics and phenomenology. For example, is it possible to develop a simple and physically motivated free-energy model that can generate accurate predictions of characteristic terminal cluster size N based on experimentally tunable variables governing monomer interactions? We address this question here by directly comparing free energy-based predictions against computer simulations for one of the most approachable and idealized cluster-forming models: the short-range attractive, long-range repulsive (SALR) pair potential.8 Once the behavior for this simple system that coarse-grains over many microscopic details of the short-range interactions, electrostatic double-layers, solvent, etc., is better understood, the goal is then to expand the framework to include more complex free energy contributions relevant for specific realizable colloidal suspensions.

First, we review the canonical a priori free-energy treatment for clustering colloidal suspensions due to Groenewold and Kegel,7,22,32 where we compare its predictions for the cluster size N against a computational survey of phases comprising compact spherical aggregates. We take great care to clarify how this elegant and frequently cited model adapts the classical nucleation theory (CNT) of non-terminating droplets (or crystals)33–35 for the SALR systems of interest by treating the latter as purely attractive reference fluids superimposed with perturbative effects due to charges on the monomers and in the suspending solvent. However, while frequently cited (and adapted for related systems, e.g., driven colloids36), this model has not been systematically scrutinized against a large “test set” of cluster phases generated by gradually varying relevant independent variables, e.g., monomer surface charge Z. By conducting tests that align with the phenomenological assumptions underlying the model (e.g., apolar solvents, low cluster density), we readily find that the analytical predictive formula derived from their model exhibits a spurious scaling for the range of stable cluster sizes observable in systems governed by SALR pair potentials.

With this knowledge in-hand, we describe and validate an alternative free energy model that quantitatively predicts the characteristic cluster size N for approximately 100 different simulated SALR systems, which comprise compact spherical aggregates in the size range 6 ≤ N ≤ 60 for wide ranges in monomer packing fraction ϕ, attraction strength βε , monomer surface Z, and solvent screening length κ−1/d (notably, even finite values outside the apolar limit). In essence, we find that a framework built on classical nucleation theory can indeed describe the thermodynamics of frustrated, finite-sized clusters provided one introduces a size-dependent enthalpic penalty of interface formation that accounts for the missing coordination bonds of “surface” particles in clusters. In justifying this approach, we also examine how the number of intracluster short-range bonds scales with size; interestingly, we find a superlinear crossover at our cluster sizes that bridges the previously established scaling regimes for very small sizes37,38 (e.g., N ≤ 9) and larger, bulk-like droplets. Surprisingly, we also demonstrate that intercluster effects need not be considered to obtain correct predictions even for rather non-dilute conditions.

To systematically test the performance of free energy models for predicting equilibrium cluster formation, it is invaluable to be able to (1) rapidly generate aggregate configurations that can be analyzed in-depth and (2) unambiguously identify relevant free energy contributions. Thus, we consider one of the simplest and most frequently used models known to form self-limiting aggregates: the short-range attractive (SA), long-range repulsive (LR) pair potential.8 The combined SALR potential can be expressed as

(1)

where β = (kBT)−1 (kB is the Boltzmann’s constant and T is temperature); x = r/d is the non-dimensionalized interparticle separation; and d is the characteristic particle diameter. We include the subscripts i and j to account for multiple particle types because we follow previous protocols17,39 and examine size-polydisperse three-component mixtures that approximate colloids with 10% size polydispersity. (This favors the formation of amorphous fluid clusters over crystalline dynamically arrested clusters.17) In this context, xi,jx − (1/2)(i + j)(Δd/d), where i (or j) = − 1, 0, 1 corresponds to small, medium, and large particles, respectively, and Δd/d is a perturbation to the particle diameter. Specifically, we study mixtures comprised of 20% small, 60% medium (characteristic size d), and 20% large particles with Δd = 0.158d.

The short-range attractions are expressed via a generalized (100-50) Lennard-Jones (LJ) model,

(2)

where βε is the reference monomer-monomer attraction strength and Δε = 0.25kBT is an energetic perturbation to promote mixing of the polydisperse particles. Given its simplicity, the contribution of Eq. (2) (similar to the contact attractions in the free energy model of Groenewold and Kegel7) does not specify the microscopic or chemical details; i.e., whether the attractions arise from depletion or other short-range interactions. Generally, the range of the attraction well is approximately 0.10d.

Long-range repulsions are calculated on the basis of the repulsive portion of the Derjaguin-Landau-Verwey-Overbeek (DLVO) potential,2,3 which approximately captures the interactions of electrostatic double-layers formed around each monomer due to (homogeneously distributed) surface charge Z. This is expressed6 as

(3)

with

(4)

where βAMAX is the maximum electrostatic barrier between particles at contact, κ−1/d is the Debye-Hückel screening length, Z is the total surface charge per monomer, and λB/d is the Bjerrum length of the solvent. Crucially, this formulation neglects any long-range multi-body interactions40,41 and any charge renormalization due to counterion condensation42–45 or close monomer association.18,29 As our goal here is to test how even the simplest clustering systems might be described from a free energy perspective, we reserve incorporation of these phenomena for future studies.

In using this model, we set the average monomer packing fraction ϕ = (π/6) ρd3 (where ρd3 is number density), charge Z, and screening length κ−1/d and then independently tune the attraction strength βε to drive aggregation as if varying the amount of non-interacting depletant. In terms of the experimental control that one has over repulsive contributions to interactions in such systems, this picture is somewhat idealized: to wit, tunable repulsion-controlling parameters are more realistically (though still ignoring some possible interdependence) charge Z, solvent relative permittivity ϵR, and solvent ionic strength I. This is because even approximately, the screening length κ1/d=ϵ0ϵRkBT/(2d2NAe2I) and λB/d = e2/(4dπϵ0ϵRkBT), where ϵ0 is the vacuum permittivity, NA is the Avogadro’s number, and e is the elementary charge. For simplicity, however, we universally fix the relative Bjerrum length λB/d, which means electrostatic effects are set via combinations of Z and κ−1/d. With this experimental picture in mind, we also note that the repulsive strength in Eq. (4) can equivalently be written as βAMAX=πdϵ0ϵRΨ02/(kBT), where Ψ0 is the surface potential on the monomer (often assumed to approximately equal the ζ-potential measured via electrophoresis).

FIG. 1.

Maximum repulsion strength βAMAX = Z2(λB/d)/[1.0 + 0.5/(κ−1/d)]2 plotted as a function of surface charge Z and screening length κ−1/d, where the left and right y-axes show Z-values referenced against two different reference Bjerrum lengths λB/d. The two reference Bjerrum lengths are λB/d = 0.014, which corresponds in real units to, e.g., d = 50 nm monomers in a solvent with dielectric constant ϵR = 80 (equivalently, d = 100 nm and ϵR = 40, or d = 200 nm and ϵR = 20); and λB/d = 0.0014, which corresponds to d = 500 nm and ϵR = 80 (equivalently, d = 1 μm and ϵR = 40, or d = 2 μm and ϵR = 20). Symbols mark Z-(κ−1/d) combinations tested via simulations, where Table I lists the specific combinations tested at each packing fraction ϕ. Throughout the manuscript, simulations are referenced by the Z-values on the left y-axis, i.e., Z = 3, 4, 6, 8, 10, 12, and 15. Contours mark βAMAX = 0.10, 0.50, 1.0, and 2.0 from bottom to top.

FIG. 1.

Maximum repulsion strength βAMAX = Z2(λB/d)/[1.0 + 0.5/(κ−1/d)]2 plotted as a function of surface charge Z and screening length κ−1/d, where the left and right y-axes show Z-values referenced against two different reference Bjerrum lengths λB/d. The two reference Bjerrum lengths are λB/d = 0.014, which corresponds in real units to, e.g., d = 50 nm monomers in a solvent with dielectric constant ϵR = 80 (equivalently, d = 100 nm and ϵR = 40, or d = 200 nm and ϵR = 20); and λB/d = 0.0014, which corresponds to d = 500 nm and ϵR = 80 (equivalently, d = 1 μm and ϵR = 40, or d = 2 μm and ϵR = 20). Symbols mark Z-(κ−1/d) combinations tested via simulations, where Table I lists the specific combinations tested at each packing fraction ϕ. Throughout the manuscript, simulations are referenced by the Z-values on the left y-axis, i.e., Z = 3, 4, 6, 8, 10, 12, and 15. Contours mark βAMAX = 0.10, 0.50, 1.0, and 2.0 from bottom to top.

Close modal

As illustrated in Fig. 1, we conduct a wide survey of Z-κ−1/d combinations designed to span the weakest repulsions that produce self-limiting aggregates (i.e., near the boundary of macrophase separation) to repulsions with strengths up to AMAX ≈ 2.0kBT. Here, note that to examine this range of repulsion strengths referenced against any plausible relative Bjerrum length λB/d (e.g., λB/d = 0.014, corresponding to d = 50 nm monomers suspended in room temperature water with λB = 0.7 nm), one must consider monomers with very low effective charge density. Throughout the publication, we reference Z-values based on the choice of λB/d = 0.014, though choosing a different reference λB/d simply renormalizes the range of Z under consideration, with an example of this rescaling given in Fig. 1. All of the parameter combinations (ϕ, Z, κ−1/d) we examine are listed in Table I by their respective critical attraction strengths (discussed below). Finally, note that throughout the remainder of the publication, we notate βui,jSALR(xi,j) as βu(r) for aesthetic simplicity unless otherwise indicated.

We generate the configurations of cluster phases via three-dimensional (3D) MD simulations of the ternary SALR mixtures described above, where we generate trajectories using LAMMPS.46 We perform simulations in the NVT ensemble with periodic boundary conditions using an integration time step of dt=0.001d2m/(kBT) (taking the mass m = 1) and fix temperature via a Nosé-Hoover thermostat with time-constant τ = 2000dt. As outlined in Table I, we consider many combinations of charge Z and screening length κ−1/d at four different packing fractions: ϕ = 0.015, 0.030, 0.060, and 0.120 (where we simulate Nbox = 1920, 2960, 6800, and 6800 particles, respectively). Beginning with randomized initial configurations, we equilibrate systems at ϕ = 0.015, 0.030, 0.060, and 0.120 for 3 × 107, 1 × 107, 3 × 106, and 2 × 106 steps, respectively, and confirm that they are equilibrated on the basis of energy convergence and visualization, where the latter shows that the systems are ergodic (aggregates undergo frequent intra- and intercluster rearrangements and exchanges). We cut off the pair potential for a given Z and κ−1/d such that the interaction strength at distance xi,jc (note the explicit use of the mixture notation) is βui,j(xi,jc)2e3 and the force is simultaneously d[βui,j(xi,jc)]/dxi,j1e3.

TABLE I.

Critical attraction strengths βε determined from MD simulations at various ϕ as a function of surface charge Z and screening length κ−1/d. Conditions with listed βε values are those used for our analysis and discussion. Symbols below the Z values correspond to those used in Figs. 2-7 (symbols are kept constant for various κ−1/d). Note that maximum repulsion strengths βAMAX (see Eq. (4)) are calculated based on a reference relative Bjerrum length of λB/d = 0.014.

Z
3468101215
κ−1/d
ϕ = 0.015(blue) 0.7 … … … … … … 6.55 
0.8 … … … … … … 6.80 
1.0 … … … 5.55 6.00 6.40 7.10 
1.2 … … … 5.65 6.10 … … 
1.5 … … 5.35 5.80 6.30 6.80 … 
2.0 … 5.05 5.50 5.95 6.45 7.00 7.90 
2.5 … … 5.55 6.00 6.60 … … 
3.0 … 5.10 5.55 6.05 6.60 … … 
4.0 4.95 5.10 5.60 6.10 6.65 … … 
ϕ = 0.030(yellow) 0.7 … … … … … … 6.30 
0.8 … … … … … … … 
1.0 … … … 5.30 5.70 6.15 6.75 
1.2 … … … 5.45 5.80 … … 
1.5 … … 5.15 5.55 5.95 6.45 … 
2.0 … 4.80 5.20 5.65 6.10 6.55 7.25 
2.5 … … 5.20 5.70 6.20 … … 
3.0 … 4.90 5.25 5.70 6.20 … … 
4.0 4.70 4.90 5.30 5.70 6.20 … … 
ϕ = 0.060(orange) 0.7 … … … … … … 6.00 
0.8 … … … … … … … 
1.0 … … … 5.00 5.40 5.65 6.25 
1.2 … … 4.75 5.10 5.45 … … 
1.5 … … 4.80 5.15 5.50 5.80 … 
2.0 … 4.55 4.85 5.20 5.55 5.80 6.40 
2.5 … … 4.90 5.20 5.60 … … 
3.0 … 4.60 4.85 … 5.60 … … 
4.0 4.40 4.60 4.85 5.20 5.60 … … 
ϕ = 0.120(red) 0.7 … … … … … … 5.20 
0.8 … … … … … … 5.20 
1.0 … … … … … 4.95 5.20 
1.2 … … … … … … … 
1.5 … … … … 4.75 4.95 5.20 
2.0 … … … 4.60 4.75 4.95 … 
2.5 … … … 4.60 … … … 
3.0 … … … … … … … 
4.0 … … … … … … … 
Z
3468101215
κ−1/d
ϕ = 0.015(blue) 0.7 … … … … … … 6.55 
0.8 … … … … … … 6.80 
1.0 … … … 5.55 6.00 6.40 7.10 
1.2 … … … 5.65 6.10 … … 
1.5 … … 5.35 5.80 6.30 6.80 … 
2.0 … 5.05 5.50 5.95 6.45 7.00 7.90 
2.5 … … 5.55 6.00 6.60 … … 
3.0 … 5.10 5.55 6.05 6.60 … … 
4.0 4.95 5.10 5.60 6.10 6.65 … … 
ϕ = 0.030(yellow) 0.7 … … … … … … 6.30 
0.8 … … … … … … … 
1.0 … … … 5.30 5.70 6.15 6.75 
1.2 … … … 5.45 5.80 … … 
1.5 … … 5.15 5.55 5.95 6.45 … 
2.0 … 4.80 5.20 5.65 6.10 6.55 7.25 
2.5 … … 5.20 5.70 6.20 … … 
3.0 … 4.90 5.25 5.70 6.20 … … 
4.0 4.70 4.90 5.30 5.70 6.20 … … 
ϕ = 0.060(orange) 0.7 … … … … … … 6.00 
0.8 … … … … … … … 
1.0 … … … 5.00 5.40 5.65 6.25 
1.2 … … 4.75 5.10 5.45 … … 
1.5 … … 4.80 5.15 5.50 5.80 … 
2.0 … 4.55 4.85 5.20 5.55 5.80 6.40 
2.5 … … 4.90 5.20 5.60 … … 
3.0 … 4.60 4.85 … 5.60 … … 
4.0 4.40 4.60 4.85 5.20 5.60 … … 
ϕ = 0.120(red) 0.7 … … … … … … 5.20 
0.8 … … … … … … 5.20 
1.0 … … … … … 4.95 5.20 
1.2 … … … … … … … 
1.5 … … … … 4.75 4.95 5.20 
2.0 … … … 4.60 4.75 4.95 … 
2.5 … … … 4.60 … … … 
3.0 … … … … … … … 
4.0 … … … … … … … 

To make our analysis tractable, we focus on states corresponding with the onset of clustering (i.e., at the cluster transition locus), where the phases are composed of fluid aggregates with characteristic size N, but the systems have not yet begun to form percolated phases or become dynamically arrested. To characterize the size of equilibrium aggregates, we calculate cluster-size distributions (CSDs), which quantify the probability p(N) of observing clusters comprising N particles. Here, we follow the established convention8,14,15,17 of considering two monomers as part of the same cluster if they are directly bonded to one another (i.e., within the range of the attractive well) or each directly bonded to a shared neighbor (i.e., are connected via some percolating pathway).

In turn, to locate the cluster transition locus, we make sweeps in attraction strength βε (at increments of Δε = 0.05kBT) and identify states at the onset of clustering based on the following criteria: (1) the p(N) distribution exhibits a visibly apparent local maximum (mode) at some 1 < NNbox, where the corresponding local minimum between N = 1 and N is notated as Nmin; and (2) that 80% of the particles in the system participate in aggregates of size NNmin, i.e., 0.80=n=NminNboxp(N), where p(N) is appropriately normalized. Taken together, these conditions correspond to the emergence of meaningful bimodality (coexistence) in p(N) between N = 1 and the cluster mode N. In this way, we obtain the characteristic cluster size N associated with a particular combination of ϕ, Z, and κ−1/d and the corresponding critical attraction strength βε. All of the parameter combinations we consider in our analysis are listed by their respective βε values in Table I.

Before discussing free energy models for characteristic cluster size N, we begin by briefly describing the cluster morphologies under examination: for the approximately 100 different combinations of packing fraction ϕ, surface charge Z, and screening length κ−1/d that we consider (listed in Table I), we observe phases at the corresponding critical attraction strengths βε that comprise compact spherical clusters with characteristic sizes in the range 6 ≤ N ≤ 60, as plotted in Fig. 2. In terms of cluster shape, we find that by measuring the radius of gyration RG/d and plotting it versus cluster size N, our results obey the relation

(5)

where α(ϕ) is a ϕ-dependent prefactor of magnitude approximately 1/2 (hereafter notated α). Together with the fractal dimension df = 3, this signifies that the aggregates are compact objects, and visual inspection of the MD trajectories confirms that the clusters are indeed highly packed amorphous droplets that are spherical on average and undergo frequent intracluster rearrangement and intercluster exchange (seen previously in Refs. 17 and 39). As shown in the inset of Fig. 2, the clusters do become slightly less packed with increasing ϕ, which is attributable to an increasing frequency of intercluster exchange. (These transfer events tend to instantaneously but, on average, isotropically distort the clusters, effectively expanding them.) We discuss trends in the cluster size and shape from a different perspective (and in more detail) in Paper I.63 

FIG. 2.

(a) Measured cluster size N versus screening length κ−1/d for all ϕ, Z, and κ−1/d combinations tested. Blue, yellow, orange, and red symbols correspond to measurements from simulations at ϕ = 0.015, 0.030, 0.060, and 0.120, respectively. Contours are guides to the eye for constant Z: from top to bottom, Z = 3.0 (no line), 4.0, 6.0, 8.0, 10.0, 12.0, and 15.0. These contours are plotted according to the formula N/Nest=1.0+1.5/(κ1/d)2, where Nest is the estimated cluster size in the Coulombic limit (i.e., κ−1/d → ∞). (b) Cluster radius of gyration RG/d versus characteristic cluster size N, both measured from MD simulations. Lines are empirical fits of the form RG/d = αN∗1/3, where α is a dimensionless prefactor corresponding to α = 0.45, 0.49, 0.53, and 0.60 for ϕ = 0.015, 0.030, 0.060, and 0.120, respectively. Symbol types in (a) and (b) correspond to constant charge Z as listed in Table I (note that we test various screening lengths κ−1/d at each Z).

FIG. 2.

(a) Measured cluster size N versus screening length κ−1/d for all ϕ, Z, and κ−1/d combinations tested. Blue, yellow, orange, and red symbols correspond to measurements from simulations at ϕ = 0.015, 0.030, 0.060, and 0.120, respectively. Contours are guides to the eye for constant Z: from top to bottom, Z = 3.0 (no line), 4.0, 6.0, 8.0, 10.0, 12.0, and 15.0. These contours are plotted according to the formula N/Nest=1.0+1.5/(κ1/d)2, where Nest is the estimated cluster size in the Coulombic limit (i.e., κ−1/d → ∞). (b) Cluster radius of gyration RG/d versus characteristic cluster size N, both measured from MD simulations. Lines are empirical fits of the form RG/d = αN∗1/3, where α is a dimensionless prefactor corresponding to α = 0.45, 0.49, 0.53, and 0.60 for ϕ = 0.015, 0.030, 0.060, and 0.120, respectively. Symbol types in (a) and (b) correspond to constant charge Z as listed in Table I (note that we test various screening lengths κ−1/d at each Z).

Close modal

In terms of cluster number size N, there are two important observations from Fig. 2: (1) characteristic cluster size depends only weakly on the packing fraction for the range of 0.015 ≤ ϕ ≤ 0.120; and (2) the morphologies associated with unscreened electrostatic repulsions (i.e., κ−1/d → ∞) are effectively generated when the screening length approaches κ−1/d ≈ 4.0. As shown by considering Figs. 2(a) and 2(b) simultaneously, increasing the packing fraction ϕ (given fixed Z and κ−1/d) does not systematically shift N but does slightly inflate the cluster radius RG/d. (We do note that the CSD peaks at N also become wider with increasing ϕ due to more frequent intercluster contacts.) The second point is apparent based on Fig. 2(a), which demonstrates that for the larger screening lengths κ−1/d tested, cluster sizes N at fixed ϕ and Z have already nearly reached asymptotic values, i.e.,

(6)

This ability to access the Coulombic limit at finite κ−1/d is important for Secs. III B and III C.

We now begin our discussion of the canonical framework for cluster formation due to Groenewold and Kegel7 (with subsequent follow-ups22,32), with an emphasis on making clear important concepts and assumptions underpinning the model. The model aims to predict characteristic cluster size N for large and perfectly monodisperse aggregates governed by short-range attractions (SA) and long-range (LR) unscreened Coulombic interactions between monomers (the subscript alludes to the κ−1/d → ∞ limit). This prediction necessarily begins with an expression for the extensive free energy βΔF of cluster formation as a function of N (agnostic to N),

(7)

where βFN and βF1 are the free energies of the N-sized clusters and monomers, respectively.

The free energy change βΔF is broken into reference and perturbative contributions: the reference portion is taken to be the free energy of aggregate formation for a SA (i.e., purely attractive) fluid, which can be described via the classical nucleation theory (CNT) for large droplets (or crystals).33–35 Meanwhile, the perturbations are any contributions to the free energy due to the electrostatic effects. This is simply expressed as

(8)

where we detail these (reference) attractive and (perturbative) repulsive free energy differentials in order below.

The CNT-based free energy contributions of the reference SA system comprise two terms, which capture competing effects that scale with aggregate volume and surface area, respectively. The first term accounts for the transfer of monomers from the low-density dispersed phase to the dense (bulk) fluid or crystal phase corresponding to the cluster interior. This transfer is characterized by a favorable change in chemical potential per particle with the magnitude βΔμ0SA. The second term is an enthalpic penalty47,48 characterized by surface tension βγSAd2, which accounts for the relative number of “missing” intracluster coordination bonds zc,m of the particles at the droplet surface relative to, e.g., the bulk-like coordination number zc,0 of the cluster interior.49 These contributions can be written as

(9)

where, reflecting our observed morphologies, we incorporate the expression for cluster surface area assuming spherical droplets with radius Rc/d. Going forward, this radius is considered interchangeable with the radius of gyration within some O(1) prefactor, i.e., RcRG.

In turn, the perturbative electrostatic contributions are treated as arising from unscreened repulsions acting between all intracluster pairs of particles (i.e., N(N − 1)/2 ≈ N2/2 interactions), which can be written as

(10)

where 〈βuLR〉 ≈ Z2(λB/d)/(Rc/d) is the Coulombic limit (κ−1/d → ∞) of the DLVO-type potential of Eqs. (3) and (4) evaluated at r = Rc/d, which assumes that the characteristic (average) intracluster pair distance is simply the cluster radius.50 The form of Eq. (10) implies that the repulsive free-energy contribution of each monomer in the dispersed phase is truly negligible compared to the intracluster contribution, which is consistent with the choice of Groenewold and Kegel to ignore intercluster interactions, i.e., consider the limit of very low ϕ. Note that Groenewold and Kegel also originally include a term (see Eq. 18 in Ref. 7) that roughly accounts for counterion condensation,42–45 which could occur for strong bare surface charges. However, we neglect this contribution because their approximation naturally drops out of the subsequent analysis, and the coarse-grained SALR potential considered here only captures a constant net-effective charge.

Given these expressions for the free energy contributions, one can proceed to the crux of the analysis: identifying the characteristic cluster size N at which the driving force to associate per monomer is at its largest magnitude (or energetic minimum), i.e., βΔf(N)minN[βΔf(N)] where βΔf(N) ≡ βΔF(N)/N. Of course, here one requires a βΔf(N) function where the sole dependent variable is N. By combining Eqs. (9) and (10) with the known relation between cluster radius and number size RG/d = αN1/3 for compact spherical aggregates, one can readily write

(11)

and evaluate its derivative to find the global minimum

(12)

which, dropping prefactors, gives the scaling relation

(13)

This states that the cluster size is simply governed by the strength of the surface energy relative to the characteristic strength of electrostatic repulsion.

To write Eq. (13) completely in terms of experimentally tunable parameters, one then approximates22,47,48 the surface tension of the SA reference fluid βγSAd2 as scaling like the attraction strength βε multiplied by the aforementioned number of missing bonds per surface particle zc,m (divided by a “surface area” per monomer Am), i.e.,

(14)

Because zc,m is considered constant with respect to N for large, low-curvature droplets, combining Eqs. (13) and (14) leads to the master a priori scaling relation,

(15)

Reintroducing prefactors, Eq. (15) is written as N=αν0βε/[Z2(λB/d)], where α remains from the repulsive term in Eq. (11) and ν0 is a prefactor that is the product of zc,m and some conversion factor to arrive at a surface energy per area.

Given our wide survey of compact spherical cluster morphologies, we can perform the first systematic test of the master scaling law given by Eq. (15) for SALR pair potentials by plotting measured cluster sizes N for systems with sufficiently large screening lengths κ−1/d at various ϕ, Z, and βε. Specifically, in Fig. 3, we plot N values observed at critical attraction strengths βε and screening length κ−1/d = 4.0, where the latter corresponds to effectively unscreened systems (see Section III A) as assumed in writing Eq. (15). Here, we note that we use the version of Eq. (15) that incorporates prefactors α and ν0, which shift predicted sizes approximately in line with the measured N values (of course, including or excluding these prefactors does not affect scaling itself).

FIG. 3.

(a) Measured cluster size N in the Coulombic limit (approximated by systems with κ−1/d = 4.0) versus the master scaling ratio of Eq. (15) plotted using measured critical attraction strengths βε and corresponding characteristic repulsion strengths Z2(λB/d). Blue, yellow, and orange symbols correspond to measurements from simulations at ϕ = 0.015, 0.030, and 0.060, respectively, for charges Z = 3.0, 4.0, 6.0, 8.0, and 10.0 (top to bottom). Thick black line corresponds to the empirical scaling of Eq. (16) with exponent of m = 3/4 (i.e., N{βε/[Z2(λB/d)]}3/4) and dark (light) purple shadings correspond to 10% (20%) deviation from this scaling. Thin black lines show scalings for alternate exponents, where the m = 1 scaling (see Eq. (15)) derives from the canonical free energy model of Groenewold and Kegel.7,22,32 Note that in this figure, we plot predicted cluster sizes (x-axis) based on including the ϕ-dependent prefactor α for the radius of gyration (see Fig. 2) and the (here, arbitrary) constant prefactor ν0 ≈ 3.40 (see text). Symbol types correspond to constant charge Z as listed in Table I (note that we test various screening lengths κ−1/d at each Z).

FIG. 3.

(a) Measured cluster size N in the Coulombic limit (approximated by systems with κ−1/d = 4.0) versus the master scaling ratio of Eq. (15) plotted using measured critical attraction strengths βε and corresponding characteristic repulsion strengths Z2(λB/d). Blue, yellow, and orange symbols correspond to measurements from simulations at ϕ = 0.015, 0.030, and 0.060, respectively, for charges Z = 3.0, 4.0, 6.0, 8.0, and 10.0 (top to bottom). Thick black line corresponds to the empirical scaling of Eq. (16) with exponent of m = 3/4 (i.e., N{βε/[Z2(λB/d)]}3/4) and dark (light) purple shadings correspond to 10% (20%) deviation from this scaling. Thin black lines show scalings for alternate exponents, where the m = 1 scaling (see Eq. (15)) derives from the canonical free energy model of Groenewold and Kegel.7,22,32 Note that in this figure, we plot predicted cluster sizes (x-axis) based on including the ϕ-dependent prefactor α for the radius of gyration (see Fig. 2) and the (here, arbitrary) constant prefactor ν0 ≈ 3.40 (see text). Symbol types correspond to constant charge Z as listed in Table I (note that we test various screening lengths κ−1/d at each Z).

Close modal

In Fig. 3, we do indeed observe a master ϕ-independent relation between the N values measured in simulations and the relative strength of attractions and repulsions between monomers, i.e., the ratio βε/[Z2(λB/d)]; however, the observed scaling does not reflect the exponent of 1 that is expected based on the free energy model underlying Eq. (15). Instead, we clearly observe the empirical relation,

(16)

for various cluster sizes and packing fractions. This immediately begs the questions: what alternative (and, ideally, comparatively simple) free energy model for SALR systems results in this softer master scaling? Furthermore, can this alternative model also readily predict N for finite screening lengths κ−1/d?

To ascertain what new model can capture the empirically observed scaling in Fig. 3 (and be extended for generic κ−1/d), we first ought to identify which of the current free energy terms in Eqs. (9) and (10) correctly (or incorrectly) describe the energetics of cluster formation in the MD simulations. Given its simplicity, the most straightforward candidate to consider is the repulsive free energy contribution of Eq. (10), which we can test against MD configurations by adding up the total repulsive energies (between all intracluster pairs of monomers) of simulated clusters as a function of characteristic size N.

FIG. 4.

(a) Total intracluster repulsion energy βΔFLR scaled by maximum repulsion barrier βAMAX = Z2(λB/d)/[1.0 + 0.5/(κ−1/d)]2 and ϕ-dependent prefactor α for the radius of gyration (see Fig. 2) plotted versus cluster size N for Z = 3.0, 4.0, 6.0, 8.0, and 10.0 (top to bottom) and κ−1/d = 4.0 (effectively κ−1/d → ∞). Blue, yellow, and orange symbols correspond to measurements from simulations at ϕ = 0.015, 0.030, and 0.060, respectively. (b) Same but for κ−1/d = 4.0, 3.0, 2.0, and 1.0 at all correspondingly tested Z values (see Table I). For (a) and (b), thick black line corresponds to the expression βΔFLR/[βAMAX/(2α)] = N∗5/3 and dark (light) purple shadings correspond to 10% (20%) deviation from this scaling. Symbol types in (a) and (b) correspond to constant charge Z as listed in Table I (note that we test various screening lengths κ−1/d at each Z).

FIG. 4.

(a) Total intracluster repulsion energy βΔFLR scaled by maximum repulsion barrier βAMAX = Z2(λB/d)/[1.0 + 0.5/(κ−1/d)]2 and ϕ-dependent prefactor α for the radius of gyration (see Fig. 2) plotted versus cluster size N for Z = 3.0, 4.0, 6.0, 8.0, and 10.0 (top to bottom) and κ−1/d = 4.0 (effectively κ−1/d → ∞). Blue, yellow, and orange symbols correspond to measurements from simulations at ϕ = 0.015, 0.030, and 0.060, respectively. (b) Same but for κ−1/d = 4.0, 3.0, 2.0, and 1.0 at all correspondingly tested Z values (see Table I). For (a) and (b), thick black line corresponds to the expression βΔFLR/[βAMAX/(2α)] = N∗5/3 and dark (light) purple shadings correspond to 10% (20%) deviation from this scaling. Symbol types in (a) and (b) correspond to constant charge Z as listed in Table I (note that we test various screening lengths κ−1/d at each Z).

Close modal

As shown in Fig. 4, we observe that the repulsive free energy contribution of Eq. (10) quantitatively describes MD results in the unscreened limit and, with a simple extension, also works for finite screening lengths κ−1/d; in other words, the current perturbative free energy term capturing electrostatics is self-consistent and should be retained. In Fig. 4(a), we see that βΔFLR measured in simulations, when normalized by the maximum repulsion barrier βAMAX = Z2(λB/d) (corresponding to the κ−1/d → ∞ limit of Eq. (4)), scales as N5/3. Of course, this N5/3 scaling is expected given N2 intracluster pair interactions occurring on the lengthscale of the cluster radius, which scales as N1/3 (see Eq. (10)). Meanwhile, Fig. 4(b) demonstrates that the same scaling holds for finite κ−1/d away from the Coulombic limit provided one appeals to the more generalized form of Eq. (4) for the maximum repulsive barrier energy, i.e., βAMAX = Z2(λB/d)/[1.0 + 0.5/(κ−1/d)]2.

Given that intracluster repulsions scale as expected (with N5/3), the simplest extensive free energy expression (resembling that of Groenewold and Kegel) that readily leads to the empirically observed scaling in Fig. 3 is one where the surface-energy penalty, rather than scaling as N2/3, instead effectively scales with a lesser exponent,

(17)

Here, ν1 is some (as yet undetermined) dimensionless prefactor distinct from the ν0 above. In turn, it is easily shown that solving Eq. (17) for βΔf(N)minN[βΔf(N)] results in the generalized scaling N ∝ {βε/[βAMAX]}3/4 or, in the unscreened limit, N{βε/[Z2(λB/d)]}3/4.

During the remainder of this section, our ultimate goal is to demonstrate that this reduced exponent for the surface energy term naturally emerges for our clustered systems because the effective energy penalty is dependent on cluster size N in the range 6 ≤ N ≤ 60. Conceptually, this size-dependence for the surface energy echoes the long-established notion that the generalized surface-tension of a liquid droplet with high curvature γ(R) will depart from the reference surface tension γ of a planar liquid-vapor interface (or very large droplet with low curvature). Indeed, starting with the pioneering work by Tolman,51 a vast number of studies have been dedicated to measuring first- and/or second-order corrections for γ(R)/γ (the classic first order correction depends on the “Tolman length”) to better model, e.g., homogeneous nucleation, but this topic continues to be an active and challenging area of research even for model systems like the LJ fluid.52–58 Compared to these studies, which are especially difficult given their general focus on critically unstable droplet formation (usually droplets with radius R ≈ 4d at the smallest), the following analysis is notable because we consider stable droplets with effective surface tensions dominated by short-range attractive bonds (much shorter than, e.g., the LJ attraction range) and radii of less than three particle diameters.

Specifically, to capture this size-dependent surface energy, one ought to account for an N-dependent number of missing co-ordination bonds zc,m(N) for the surface particles relative to the reference bulk (interior) co-ordination number zc,0. The surface energy penalty in Eq. (17) can then be written as

(18)

with the (to be demonstrated) scaling

(19)

where we still assume that the number of surface particles at least roughly scales as N2/3, i.e., proportional to the squared cluster radius (RG/d)2 = α2N2/3, though making a formal distinction between interior and surface particles is difficult for small N (as discussed later). To demonstrate that the scaling in Eq. (19) is reasonable, we show in Figs. 5 and 6 that this size-dependence for zc,m(N) = zc,0zc(N) originates based on the coordination number of (surface) particles zc(N) measured from MD configurations, which we calculate from the extensive number of intracluster bonds nB(N). Given our measurement of nB(N) is at the root of much of this analysis, we consider its behavior first and proceed backwards to the scaling of Eq. (19).

Looking towards estimating zc,m(N), consider in Fig. 5(a) the extensive number of intracluster bonds nB(N) measured from MD simulations, where we observe a superlinear growth rate over the range of cluster sizes that we generate. Interestingly, this superlinear behavior contrasts with known small- and large-cluster limits, which are linear in N. Here, nB(N) is nicely captured at each packing fraction for 6 ≤ N ≤ 60 by the empirical expression,

(20)

where k is a ϕ-specific O(1) prefactor59 and we include a division by 2 for aesthetic purposes. This superlinear regime contrasts with the small cluster regime (3 ≤ N ≤ 9), where it is known37,38 that colloidal clusters dominated by SA bonds maximize their extensive bonding number according to the expression nB(N) = 3N − 6. Likewise, in the limit of large droplets, the number of bonds must scale increasingly like in the corresponding bulk fluid, i.e., nB(N) → (zbulk/2) N, where zbulk is the co-ordination number of the reference fluid (or crystal) phase.

FIG. 5.

(a) Extensive number of intracluster bonds nB(N) versus cluster size N. Blue, yellow, and orange symbols correspond to measurements from simulations at ϕ = 0.015, 0.030, and 0.060, respectively. Symbol types correspond to constant charge Z as listed in Table I (note that we test various screening lengths κ−1/d at each Z). Blue, yellow, and orange solid lines are of the empirical form nB(N) = (k/2) Nln(N) found to apply between 6 ≤ N ≤ 60, where k = 2.20, 1.95, and 1.70 with respect to ϕ. Purple line corresponds to small cluster limit37,38nB(N) = 3N − 6, which is accurate for 3 ≤ N ≤ 9. Black line corresponds to large droplet (bulk) limit nB(N) = (zbulk/2) N where we choose zbulk = 12 (see text); this limit becomes near-quantitative for dense droplets of NO(1000). Dashed blue curve is a schematic extension to the solid blue line between 60 ≤ N ≤ 500. (b) Average co-ordination number zc(N) versus cluster size N. Symbols and lines have same meaning as in (a), where the latter are calculated via the formula zc(N) = 2nB(N)/N.

FIG. 5.

(a) Extensive number of intracluster bonds nB(N) versus cluster size N. Blue, yellow, and orange symbols correspond to measurements from simulations at ϕ = 0.015, 0.030, and 0.060, respectively. Symbol types correspond to constant charge Z as listed in Table I (note that we test various screening lengths κ−1/d at each Z). Blue, yellow, and orange solid lines are of the empirical form nB(N) = (k/2) Nln(N) found to apply between 6 ≤ N ≤ 60, where k = 2.20, 1.95, and 1.70 with respect to ϕ. Purple line corresponds to small cluster limit37,38nB(N) = 3N − 6, which is accurate for 3 ≤ N ≤ 9. Black line corresponds to large droplet (bulk) limit nB(N) = (zbulk/2) N where we choose zbulk = 12 (see text); this limit becomes near-quantitative for dense droplets of NO(1000). Dashed blue curve is a schematic extension to the solid blue line between 60 ≤ N ≤ 500. (b) Average co-ordination number zc(N) versus cluster size N. Symbols and lines have same meaning as in (a), where the latter are calculated via the formula zc(N) = 2nB(N)/N.

Close modal

To quickly understand why nB(N) growth should be superlinear over this size range, we show in Fig. 5(a) extensions of the small- and large-cluster linear regimes (to large and small N where they should, respectively, fail) to demonstrate that the function nB(N) = (k/2) Nln(N) connects these otherwise disparate limits, while quantitatively overlapping with the upper reaches of the small cluster trend at N ≈ 10. To wit, notice that the characteristic slope of the small-N regime, (m = 3), differs meaningfully from the typical slope in the large-N regime of a very dense bulk fluid or crystal, which we estimate as m = zbulk/2 = 6 with zbulk = 12 because it is the sphere kissing number in three dimensions60 (this is justified later). Thus, provided zbulk is decidedly larger than 6, a superlinear regime allows for a smooth continuous growth in nB(N) with respect to N.

This connectivity between very small and large cluster sizes is clearly echoed by the next necessary quantity we must calculate: the average co-ordination number zc(N) = 2nB(N)/N = kln(N), which we show in Fig. 5(b) for the same selected clustered states.61 Here, we plot zc(N) values calculated from MD configurations, which begin to bridge the gap (up to the highest cluster sizes we observe) between the highly bond-restricted regime at small N and the bulk regime at large N , where the co-ordination number approaches zc(N) → zbulk. Notably, zc(N) varies by approximately a factor of 2 over the size range of interest 6 ≤ N ≤ 60, which underlines that the conventional practice (for larger droplets) of assuming that surface effects are size-independent is problematic for these smaller aggregates.

With zc(N) in hand, we can proceed to calculate the average number of missing bonds per particle zc,m(N), which indeed collapses onto a master curve scaling as N−1/3 (shown in Fig. 6) when the magnitude of the reference (fitting) co-ordination parameter zc,0 is set—in line with measurements of cluster interiors—at values appropriate for highly packed bulk fluids. To do this, we use the expression

(21)

where the only as-yet undetermined value is zc,0, which is the co-ordination number of the reference bulk SA fluid that represents the idealized cluster interior. For our immediate purposes, we treat this parameter as tunable and verify our choices as reasonable below. As shown in Fig. 6(a), our data approximately collapse onto a master curve with characteristic N−1/3 dependence when zc,0 = 12.0, 11.5, and 10.5 for ϕ = 0.015, 0.030, and 0.060, respectively. All of these values—especially for the lowest-density case—are reflective of bulk fluids dominated by short-range attractions, especially here given that energetic gains from bonding occur within attractive wells beyond surface contact that are approximately 0.1d in width.

FIG. 6.

(a) Average number of missing bonds per particle zc,m(N) = zc,0zc(N) versus cluster size N. Blue, yellow, and orange symbols correspond to measurements from simulations at ϕ = 0.015, 0.030, and 0.060, respectively. Fitting parameter zc,0 is the co-ordination number of the reference bulk dense fluid, found to be zc,0 = 12.0, 11.5, and 10.5 with respect to ϕ. Thick black line is a scaling guideline with the form zc,m(N) = 15.5N−1/3 and dark (light) purple shadings correspond to 10% (20%) deviation from this scaling. Symbol types correspond to constant charge Z as listed in Table I (note that we test various screening lengths κ−1/d at each Z). (b) Locally averaged intracluster co-ordination number zc(r) measured at radial positions r relative to cluster center of mass for four selected cluster phases. Blue, yellow, and orange circles are for Z = 3.0 and κ−1/d = 4.0 at ϕ = 0.015, 0.030, and 0.060, respectively, where 50 < N < 60. Blue squares are for Z = 6.0 and κ−1/d = 4.0 at ϕ = 0.015, where N ≈ 20. Arrow points to inner regions of clusters, highlighting zc(r → 0) ≈ 12.

FIG. 6.

(a) Average number of missing bonds per particle zc,m(N) = zc,0zc(N) versus cluster size N. Blue, yellow, and orange symbols correspond to measurements from simulations at ϕ = 0.015, 0.030, and 0.060, respectively. Fitting parameter zc,0 is the co-ordination number of the reference bulk dense fluid, found to be zc,0 = 12.0, 11.5, and 10.5 with respect to ϕ. Thick black line is a scaling guideline with the form zc,m(N) = 15.5N−1/3 and dark (light) purple shadings correspond to 10% (20%) deviation from this scaling. Symbol types correspond to constant charge Z as listed in Table I (note that we test various screening lengths κ−1/d at each Z). (b) Locally averaged intracluster co-ordination number zc(r) measured at radial positions r relative to cluster center of mass for four selected cluster phases. Blue, yellow, and orange circles are for Z = 3.0 and κ−1/d = 4.0 at ϕ = 0.015, 0.030, and 0.060, respectively, where 50 < N < 60. Blue squares are for Z = 6.0 and κ−1/d = 4.0 at ϕ = 0.015, where N ≈ 20. Arrow points to inner regions of clusters, highlighting zc(r → 0) ≈ 12.

Close modal

In Fig. 6(b), we demonstrate that these zc,0 values are appropriate based on direct measurements of the locally averaged co-ordination number zc(r) as a function of radial position within clusters (relative to cluster center-of-mass). Here, we specifically show results from some of the largest clusters observed (50 < N < 60), which are most likely to possess bulk-like interiors as r → 0; indeed, it is evident that zc(r → 0) ≈ 12 for these larger clusters, though the limiting value (as above) slightly decreases as ϕ increases, presumably due to the previously discussed trend in intracluster density. We also observe in Fig. 6(b) that the zc(r → 0) limit is similar even for smaller clusters, e.g., N ≈ 20, where central particles can still be surrounded by a packed shell of intracluster neighbors.

Taken altogether, the results of Figs. 5 and 6 nicely justify the choice to quantify the surface energy penalty of cluster formation from the perspective of a size-dependence in the relative number of missing bonds zc,m(N). Before moving on to consider the impact of the scaling relationships in Eqs. (18) and (19) for predicting cluster size, we pause to note that in the analysis above, we approximate zc,m(N) for surface particles based on an average measurement of nB(N) for all cluster constituents. We take this somewhat imprecise approach because it draws upon relatively unambiguous measurable quantities and bypasses the fraught process of definitively distinguishing between surface and interior particles (consider, e.g., Fig. 6(b)). Our approximation is sufficient for the proof-of-concept analysis here, but we imagine a more exacting analysis in this vein would be a worthwhile future endeavor.62 

Based on the new scaling for the free energy surface penalty justified in Section III D and the generalized free energy term for repulsive contributions in Section III C, we can return to the extensive free energy model of Eq. (17) and readily derive a new master equation for predicting cluster size N based on experimentally tunable parameters,

(22)

where, as before, α is the known ϕ-dependent prefactor relating the cluster radius and number size and ν2 is a constant similar to those above that scales the surface energy penalty, which we treat as an empirical tuning parameter. Eq. (22) is the central result of this publication.

FIG. 7.

Measured cluster size N versus predicted cluster size from Eq. (22), where the latter formula is a function of ϕ-dependent radius of gyration coefficient α (see Fig. 2); critical attraction strength βε; and maximum repulsive barrier height βAMAX = Z2(λB/d)/[1.0 + 0.5/(κ−1/d)]2. The constant prefactor ν22π scales the surface energy penalty associated with aggregation (see text). Thick black line corresponds to the Eq. (22) relation and dark (light) purple shadings correspond to 10% (20%) deviation from this scaling. Blue, yellow, orange, and red symbols correspond to results from simulations at ϕ = 0.015, 0.030, 0.060, and 0.120, respectively. Symbol types correspond to constant charge Z as listed in Table I (note that we test various screening lengths κ−1/d at each Z). The three illustrated clusters are instantaneous configurations observed in MD simulations, with blue, yellow, and orange spheres corresponding to small, medium, and large particles, respectively.

FIG. 7.

Measured cluster size N versus predicted cluster size from Eq. (22), where the latter formula is a function of ϕ-dependent radius of gyration coefficient α (see Fig. 2); critical attraction strength βε; and maximum repulsive barrier height βAMAX = Z2(λB/d)/[1.0 + 0.5/(κ−1/d)]2. The constant prefactor ν22π scales the surface energy penalty associated with aggregation (see text). Thick black line corresponds to the Eq. (22) relation and dark (light) purple shadings correspond to 10% (20%) deviation from this scaling. Blue, yellow, orange, and red symbols correspond to results from simulations at ϕ = 0.015, 0.030, 0.060, and 0.120, respectively. Symbol types correspond to constant charge Z as listed in Table I (note that we test various screening lengths κ−1/d at each Z). The three illustrated clusters are instantaneous configurations observed in MD simulations, with blue, yellow, and orange spheres corresponding to small, medium, and large particles, respectively.

Close modal

As demonstrated in Fig. 7, Eq. (22) successfully predicts characteristic cluster sizes for the vast majority of our ≈100 cases over various Z-κ−1/d combinations and near order-of-magnitude ranges in both size 6 ≤ N ≤ 60 and bulk monomer packing fraction 0.015 ≤ ϕ ≤ 0.120. This wide applicability is notable as the underlying free energy framework remains very simple: to wit, intercluster effects can evidently be neglected even as conditions become less dilute (e.g., ϕ ≈ 0.120), though the current model cannot predict more subtle trends known for the SALR model39 like the growing polydispersity of aggregates with increasing ϕ. Meanwhile, the biggest deviations between measured and predicted N (larger than 20%) occur for states combining large charge (e.g., Z = 15.0) and small screening length (e.g., κ−1/d = 0.70), which results in rather non-idealized repulsions that are both strong relative to kBT and far from the Coulombic limit. Finally, note that the value for prefactor ν2 that shifts the (already collapsed) predictions into the correct range is ν22π, which we expect should apply rather generally for compact colloidal clusters as it simply converts between measurements of the cluster surface-size based on population and radius.

We have validated a new and readily applied formula (Eq. (22)) that can predict the characteristic cluster size N for idealized SALR suspensions as a function of the variables controlling monomer-monomer interactions (including attraction strength βε, surface charge Z, and screening length κ−1/d). Eq. (22) and its underlying free energy model represent a semi-empirical adaptation and extension of the canonical free energy model due to Groenewold and Kegel,7,22,32 where we find that the latter exhibits a spurious scaling of N away from the large-droplet limit with respect to the ratio of attractive and repulsive interaction strengths driving aggregation. We subsequently find that Eq. (22) performs excellently based on direct comparisons of predicted cluster sizes and measurements of N from MD simulations of approximately 100 different systems for very wide ranges in ϕ, Z, and κ−1/d, where we examine states at the onset of clustering that exhibit compact spherical aggregates in the size range 6 ≤ N ≤ 60.

The predictive quality of Eq. (22) demonstrates that a simple free energy model, which treats SALR systems as reference SA fluids (via classical nucleation theory) with additive repulsive perturbations due to electrostatic effects, can be applied down to extremely small cluster sizes (N < 10) provided one properly corrects for surface effects at small N. Conceptually, this is in the spirit of long-standing investigations regarding size-dependent surface tensions in small droplets,51–58 and practically, we find that one can treat the energy penalty of “interface” formation as a function of an N-dependent number of missing co-ordination bonds zc,m(N) for surface particles (referenced against the co-ordination number in the bulk fluid). Here, this picture is validated in part by configurational analysis of the number of extensive intracluster bonds nB(N), which exhibits a superlinear scaling regime over the size range 6 ≤ N ≤ 60. Meanwhile, based on the form of the free energy model, we confirm that intercluster effects can be neglected even for rather non-dilute conditions (e.g., ϕ = 0.120), which are reflected by our observation that cluster size N exhibits little variability with respect to ϕ given otherwise fixed conditions.

We look forward to testing the predictive capability of Eq. (22) for real colloidal suspensions that exhibit equilibrium cluster phases, which could help bolster whether SALR pair potentials are a sufficient (if idealized) description of experimental systems. For instance, there has been recent discussion16,18,29 in the literature as to whether accounting for charge renormalization during aggregation is necessary for describing cluster behavior; likewise, Groenewold and Kegel initially postulated that non-trivial charge effects42–45 could affect the free energy picture in certain limits. Of course, these effects are not captured by the canonical SALR pair potential examined here, but we now possess a free energy model known to describe this simpler system. Thus, ascertaining whether cluster size N scales in experiments similarly to the empirical scaling of Eq. (22) would help clarify the degree to which phenomenology and interpretive guidelines derived from the pairwise model are appropriate for real systems. Similarly, it is fascinating to consider how accounting for size-dependent surface effects, here so crucial for producing quantitative predictions, might change for less compact (e.g., elongated) aggregates than those considered here.

This work was partially supported by the National Science Foundation (Grant No. 1247945) and the Welch Foundation (No. F-1696). We acknowledge the Texas Advanced Computing Center (TACC) at The University of Texas at Austin for providing HPC resources.

1.
M. V.
Smoluchowski
, “
Drei vorträge über diffusion, brownsche bewegung und koagulation von kolloidteilchen
,”
Phys. Z.
17
,
557
585
(
1916
).
2.
B. V.
Derjaguin
and
L.
Landau
, “
Theory of the stability of strongly charged lyophobic sols and of the adhesion of strongly charged particles in solution of electrolytes
,”
Acta Physicochim. URSS
14
,
633
662
(
1941
).
3.
E. J.
Verwey
and
J. T. G.
Overbeek
,
Theory of the Stability Lyophobic Colloids
(
Elsevier
,
New York, NY, USA
,
1948
).
4.
M. Y.
Lin
,
H. M.
Lindsay
,
D. A.
Weitz
,
R. C.
Ball
,
R.
Klein
, and
P.
Meakin
, “
Universality in colloid aggregation
,”
Nature
339
,
360
362
(
1989
).
5.
V.
Anderson
and
H. N. W.
Lekkerkerker
, “
Insights into phase transition kinetics from colloid science
,”
Nature
416
,
811
815
(
2002
).
6.
J. N.
Israelachvili
,
Intermolecular and Surface Forces
(
Academic Press
,
New York, NY, USA
,
2011
).
7.
J.
Groenewold
and
W. K.
Kegel
, “
Anomalously large equilibrium clusters of colloids
,”
J. Phys. Chem. B
105
,
11702
11709
(
2001
).
8.
F.
Sciortino
,
S.
Mossa
,
E.
Zaccarelli
, and
P.
Tartaglia
, “
Equilibrium cluster phases and low-density arrested disordered states: The role of short-range attraction and long-range repulsion
,”
Phys. Rev. Lett.
93
,
055701
(
2004
).
9.
A. J.
Archer
and
N. B.
Wilding
, “
Phase behavior of a fluid with competing attractive and repulsive interactions
,”
Phys. Rev. E
76
,
031501
(
2007
).
10.
J. C. F.
Toledano
,
F.
Sciortino
, and
E.
Zaccarelli
, “
Colloidal systems with competing interactions: From an arrested repulsive cluster phase to a gel
,”
Soft Matter
5
,
2390
2398
(
2009
).
11.
T.
Jiang
and
J.
Wu
, “
Cluster formation and bulk phase behavior of colloidal dispersions
,”
Phys. Rev. E
80
,
021401
(
2009
).
12.
J.-M.
Bomont
,
J.-L.
Bretonnet
,
D.
Costa
, and
J.-P.
Hansen
, “
Communication: Thermodynamic signatures of cluster formation in fluids with competing interactions
,”
J. Chem. Phys.
137
,
011101
(
2012
).
13.
P. D.
Godfrin
,
R.
Castaeda-Priego
,
Y.
Liu
, and
N. J.
Wagner
, “
Intermediate range order and structure in colloidal dispersions with competing interactions
,”
J. Chem. Phys.
139
,
154904
(
2013
).
14.
P. D.
Godfrin
,
N. E.
Valadez-Perez
,
R.
Castaneda-Priego
,
N. J.
Wagner
, and
Y.
Liu
, “
Generalized phase behavior of cluster formation in colloidal dispersions with competing interactions
,”
Soft Matter
10
,
5061
5071
(
2014
).
15.
E.
Mani
,
W.
Lechner
,
W. K.
Kegel
, and
P. G.
Bolhuis
, “
Equilibrium and non-equilibrium cluster phases in colloids with competing interactions
,”
Soft Matter
10
,
4479
4486
(
2014
).
16.
M. B.
Sweatman
,
R.
Fartaria
, and
L.
Lue
, “
Cluster formation in fluids with competing short-range and long-range interactions
,”
J. Chem. Phys.
140
,
124508
(
2014
).
17.
R. B.
Jadrich
,
J. A.
Bollinger
,
K. P.
Johnston
, and
T. M.
Truskett
, “
Origin and detection of microstructural clustering in fluids with spatial-range competitive interactions
,”
Phys. Rev. E
91
,
042312
(
2015
).
18.
T. D.
Nguyen
,
B. A.
Schultz
,
N. A.
Kotov
, and
S. C.
Glotzer
, “
Generic, phenomenological, on-the-fly renormalized repulsion model for self-limited organization of terminal supraparticle assemblies
,”
Proc. Natl. Acad. Sci. U. S. A.
112
,
E3161
E3168
(
2015
).
19.
Y.
Zhuang
and
P.
Charbonneau
, “
Recent advances in the theory and simulation of model colloidal microphase formers
,”
J. Phys. Chem. B
(published online).
20.
A. I.
Campbell
,
V. J.
Anderson
,
J. S.
van Duijneveldt
, and
P.
Bartlett
, “
Dynamical arrest in attractive colloids: The effect of long-range repulsion
,”
Phys. Rev. Lett.
94
,
208301
(
2005
).
21.
C. L.
Klix
,
C. P.
Royall
, and
H.
Tanaka
, “
Structural and dynamical features of multiple metastable glassy states in a colloidal system with competing interactions
,”
Phys. Rev. Lett.
104
,
165702
(
2010
).
22.
T. H.
Zhang
,
J.
Klok
,
R.
Hans Tromp
,
J.
Groenewold
, and
W. K.
Kegel
, “
Non-equilibrium cluster states in colloids with competing interactions
,”
Soft Matter
8
,
667
672
(
2012
).
23.
Y.
Xia
,
T. D.
Nguyen
,
M.
Yang
,
B.
Lee
,
A.
Santos
,
P.
Podsiadlo
,
Z.
Tang
,
S. C.
Glotzer
, and
N. A.
Kotov
, “
Self-assembly of self-limiting monodisperse supraparticles from polydisperse nanoparticles
,”
Nat. Nanotechnol.
7
,
479
(
2012
).
24.
A.
Yethiraj
and
A.
van Blaaderen
, “
A colloidal model system with an interaction tunable from hard sphere to soft and dipolar
,”
Nature
421
,
513
517
(
2003
).
25.
A.
Stradner
,
H.
Sedgwick
,
F.
Cardinaux
,
W. C. K.
Poon
,
S. U.
Egelhaaf
, and
P.
Schurtenberger
, “
Equilibrium cluster formation in concentrated protein solutions and colloids
,”
Nature
432
,
492
495
(
2004
).
26.
L.
Porcar
,
P.
Falus
,
W.-R.
Chen
,
A.
Faraone
,
E.
Fratini
,
K.
Hong
,
P.
Baglioni
, and
Y.
Liu
, “
Formation of the dynamic clusters in concentrated lysozyme protein solutions
,”
J. Phys. Chem. Lett.
1
,
126
129
(
2010
).
27.
Y.
Liu
,
L.
Porcar
,
J.
Chen
,
W.-R.
Chen
,
P.
Falus
,
A.
Faraone
,
E.
Fratini
,
K.
Hong
, and
P.
Baglioni
, “
Lysozyme protein solution with an intermediate range order structure
,”
J. Phys. Chem. B
115
,
7238
7247
(
2011
).
28.
K. P.
Johnston
,
J. A.
Maynard
,
T. M.
Truskett
,
A. U.
Borwankar
,
M. A.
Miller
,
B. K.
Wilson
,
A. K.
Dinin
,
T. A.
Khan
, and
K. J.
Kaczorowski
, “
Concentrated dispersions of equilibrium protein nanoclusters that reversibly dissociate into active monomers
,”
ACS Nano
6
,
1357
1369
(
2012
).
29.
J. I.
Park
,
T. D.
Nguyen
,
G.
de Queirós Silveira
,
J. H.
Bahng
,
S.
Srivastava
,
G.
Zhao
,
K.
Sun
,
P.
Zhang
,
S. C.
Glotzer
, and
N. A.
Kotov
, “
Terminal supraparticle assemblies from similarly charged protein molecules and nanoparticles
,”
Nat. Commun.
5
,
3593
(
2014
).
30.
E. J.
Yearley
,
P. D.
Godfrin
,
T.
Perevozchikova
,
H.
Zhang
,
P.
Falus
,
L.
Porcar
,
M.
Nagao
,
J. E.
Curtis
,
P.
Gawande
,
R.
Taing
,
I. E.
Zarraga
,
N. J.
Wagner
, and
Y.
Liu
, “
Observation of small cluster formation in concentrated monoclonal antibody solutions and its implications to solution viscosity
,”
Biophys. J.
106
,
1763
1770
(
2014
).
31.
P. D.
Godfrin
,
I. E.
Zarraga
,
J.
Zarzar
,
L.
Porcar
,
P.
Falus
,
N. J.
Wagner
, and
Y.
Liu
, “
Effect of hierarchical cluster formation on the viscosity of concentrated monoclonal antibody formulations studied by neutron scattering
,”
J. Phys. Chem. B
120
,
278
291
(
2016
).
32.
J.
Groenewold
and
W. K.
Kegel
, “
Colloidal cluster phases, gelation and nuclear matter
,”
J. Phys.: Condens. Matter
16
,
S4877
(
2004
).
33.
P. G.
Debenedetti
,
Metastable Liquids: Concepts and Principles
(
Princeton University Press
,
Princeton, NJ, USA
,
1997
).
34.
S.
Auer
and
D.
Frenkel
, “
Prediction of absolute crystal-nucleation rate in hard-sphere colloids
,”
Nature
409
,
1020
1023
(
2000
).
35.
R. P.
Sear
, “
Nucleation: Theory and applications to protein solutions and colloidal suspensions
,”
J. Phys.: Condens. Matter
19
,
033101
(
2007
).
36.
E.
Mani
and
H.
Löwen
, “
Effect of self-propulsion on equilibrium clustering
,”
Phys. Rev. E
92
,
032301
(
2015
).
37.
N.
Arkus
,
V. N.
Manoharan
, and
M. P.
Brenner
, “
Minimal energy clusters of hard spheres with short range attractions
,”
Phys. Rev. Lett.
103
,
118303
(
2009
).
38.
G.
Meng
,
N.
Arkus
,
M. P.
Brenner
, and
V. N.
Manoharan
, “
The free-energy landscape of clusters of attractive hard spheres
,”
Science
327
,
560
563
(
2010
).
39.
R. B.
Jadrich
,
J. A.
Bollinger
,
B. A.
Lindquist
, and
T. M.
Truskett
, “
Equilibrium cluster fluids: Pair interactions via inverse design
,”
Soft Matter
11
,
9342
9354
(
2015
).
40.
G.
Pandav
,
V.
Pryamitsyn
,
J.
Errington
, and
V.
Ganesan
, “
Multibody interactions, phase behavior, and clustering in nanoparticlepolyelectrolyte mixtures
,”
J. Phys. Chem. B
119
,
14536
14550
(
2015
).
41.
G.
Pandav
,
V.
Pryamitsyn
, and
V.
Ganesan
, “
Interactions and aggregation of charged nanoparticles in uncharged polymer solutions
,”
Langmuir
31
,
12328
12338
(
2015
).
42.
G. S.
Manning
, “
Counterion binding in polyelectrolyte theory
,”
Acc. Chem. Res.
12
,
443
449
(
1979
).
43.
S.
Alexander
,
P. M.
Chaikin
,
P.
Grant
,
G. J.
Morales
,
P.
Pincus
, and
D.
Hone
, “
Charge renormalization, osmotic pressure, and bulk modulus of colloidal crystals: Theory
,”
J. Chem. Phys.
80
,
5776
5781
(
1984
).
44.
G. V.
Ramanathan
, “
Counterion condensation in micellar and colloidal solutions
,”
J. Chem. Phys.
88
,
3887
3892
(
1988
).
45.
D. A. J.
Gillespie
,
J. E.
Hallett
,
O.
Elujoba
,
A. F.
Che Hamzah
,
R. M.
Richardson
, and
P.
Bartlett
, “
Counterion condensation on spheres in the salt-free limit
,”
Soft Matter
10
,
566
577
(
2014
).
46.
S.
Plimpton
, “
Fast parallel algorithms for short-range molecular dynamics
,”
J. Comput. Phys.
117
,
1
19
(
1995
).
47.
R. P.
Sear
, “
Classical nucleation theory for the nucleation of the solid phase of spherical particles with a short-ranged attraction
,”
J. Chem. Phys.
111
,
2001
2007
(
1999
).
48.
R. P.
Sear
, “
Low-temperature interface between the gas and solid phases of hard spheres with a short-ranged attraction
,”
Phys. Rev. E
59
,
6838
6841
(
1999
).
49.

In principle, the free-energy penalty also includes an entropic contribution due to the increased mobility particles might have at the droplet surface compared to the droplet interior; however, this contribution is often negligible.48 Groenewold and co-workers do not address this issue,7,22,32 but for our systems, where clusters possess fluid-like structures with frequent rearrangement between interior and exterior (and frequent intercluster exchange), we also expect this entropic differential to be small.

50.

Zhang and co-workers22 report the wrong exponent with respect to N for this term.

51.
R. C.
Tolman
, “
The effect of droplet size on surface tension
,”
J. Chem. Phys.
17
,
333
337
(
1949
).
52.
M. J. P.
Nijmeijer
,
C.
Bruin
,
A. B.
van Woerkom
,
A. F.
Bakker
, and
J. M. J.
van Leeuwen
, “
Molecular dynamics of the surface tension of a drop
,”
J. Chem. Phys.
96
,
565
576
(
1992
).
53.
R.
McGraw
and
A.
Laaksonen
, “
Scaling properties of the critical nucleus in classical and molecular-based theories of vapor-liquid nucleation
,”
Phys. Rev. Lett.
76
,
2754
2757
(
1996
).
54.
P. R.
ten Wolde
and
D.
Frenkel
, “
Computer simulation study of gas-liquid nucleation in a Lennard-Jones system
,”
J. Chem. Phys.
109
,
9901
9918
(
1998
).
55.
K.
Koga
,
X. C.
Zeng
, and
A. K.
Shchekin
, “
Validity of Tolman’s equation: How large should a droplet be?
,”
J. Chem. Phys.
109
,
4063
4070
(
1998
).
56.
A. E.
van Giessen
and
E. M.
Blokhuis
, “
Direct determination of the Tolman length from the bulk pressures of liquid drops via molecular dynamics simulations
,”
J. Chem. Phys.
131
,
164705
(
2009
).
57.
A.
Tröster
,
M.
Oettel
,
B.
Block
,
P.
Virnau
, and
K.
Binder
, “
Numerical approaches to determine the interface tension of curved interfaces from free energy calculations
,”
J. Chem. Phys.
136
,
064709
(
2012
).
58.
Ø.
Wilhelmsen
,
D.
Bedeaux
, and
D.
Reguera
, “
Tolman length and rigidity constants of the Lennard-Jones fluid
,”
J. Chem. Phys.
142
,
064706
(
2015
).
59.

Note that the prefactor k modestly decreases as ϕ increases: this occurs because, as discussed earlier, the cluster radius modestly increases with ϕ for fixed N; thus, clusters become less dense and exhibit correspondingly fewer bonds.

60.
F.
Pfender
and
G. M.
Ziegler
, “
Kissing numbers, sphere packings and some unexpected proofs
,”
Not. Am. Math. Soc.
51
,
873
883
(
2004
).
61.

The relation between co-ordination number and cluster size that we observe, zc(N) = kln(N) (with k ≈ 2), has a much stronger scaling than that of a similar relation reported by Godfrin et al.,14 which was given as zc(N) = 1.5[ln(N)]1/2 (here written in our choice of notation). We would simply note that the latter reaches an asymptotic co-ordination number of approximately 4 at very large droplet sizes, which would point to extremely non-compact clusters (even Bernal spiral motifs20 exhibit zc ≈ 5). In contrast, our expression, which is based on data from compact spherical aggregates at the onset of clustering, grows with cluster size and tends to approach the bulk co-ordination number zbulk = 12 of a dense attractive fluid in the large N limit, as in Fig. 5(b).

62.

In our approximate treatment, we suspect that at smallN, we are simultaneously: (1) underestimating the relative fraction of surface particles, which means the number of surface particles actually scales as Nm with m < 2/3 over the whole intermediate size range; and (2) overestimating the co-ordination number zc(N) of surface particles (underestimating zc,m(N)), which means that the number of missing surface bonds actually scales as zc,m(N) ∝ Nm with m > − 1/3. Because these errors in the exponents tend to cancel each other, we expect that the net effective N1/3 scaling of the surface term in Eq. (18) holds even given greater precision in the configurational analysis.

63.
J. A.
Bollinger
and
T. M.
Truskett
, “
Fluids with competing interactions. I. Decoding the structure factor to detect and characterize self-limited clustering
,”
J. Chem. Phys.
145
,
064902
(
2016
).